• No results found

On the use of iron in organic chemistry

N/A
N/A
Protected

Academic year: 2022

Share "On the use of iron in organic chemistry"

Copied!
20
0
0

Loading.... (view fulltext now)

Full text

(1)

molecules

Review

On the Use of Iron in Organic Chemistry

Arnar Guðmundsson1and Jan-E. Bäckvall1,2,*

1 Department of Organic Chemistry, Arrhenius Laboratory, Stockholm University, SE-10691 Stockholm, Sweden; arnar.gudmundsson@su.se

2 Department of Natural Sciences, Mid Sweden University, Holmgatan 10, 85179 Sundsvall, Sweden

* Correspondence: jeb@organ.su.se; Tel.:+46-08-674-71-78

Received: 2 March 2020; Accepted: 10 March 2020; Published: 16 March 2020  Abstract: Transition metal catalysis in modern organic synthesis has largely focused on noble transition metals like palladium, platinum and ruthenium. The toxicity and low abundance of these metals, however, has led to a rising focus on the development of the more sustainable base metals like iron, copper and nickel for use in catalysis. Iron is a particularly good candidate for this purpose due to its abundance, wide redox potential range, and the ease with which its properties can be tuned through the exploitation of its multiple oxidation states, electron spin states and redox potential.

This is a fact made clear by all life on Earth, where iron is used as a cornerstone in the chemistry of living processes. In this mini review, we report on the general advancements in the field of iron catalysis in organic chemistry covering addition reactions, C-H activation, cross-coupling reactions, cycloadditions, isomerization and redox reactions.

Keywords: iron; organic synthesis; C-H activation; C-C coupling

1. Introduction

Iron is the most abundant element on Earth by mass and is used ubiquitously by living organisms [1].

The ability of iron to assume many oxidation states (from −2 up to+6) coupled with its low toxicity makes it an attractive, versatile and useful catalyst in organic synthesis. The wide range of oxidation states available for iron and its ability to promote single electron transfer (SET) allows it to cover a wide range of transformations. In low oxidation states, iron becomes nucleophilic in character and takes part in reactions such as nucleophilic substitutions, reductions and cycloisomerizations. At higher oxidation states, iron behaves as a Lewis acid, activating unsaturated bonds and at very high oxidation states (+3 to +5) iron complexes can take part in C-H activation. Due to iron’s central position in the periodic table, it can have the property of both an “early” and “late” transition metal and with the many oxidation states available, any type of reaction is, in principle, within reach. Iron cations also bind strongly to many N- and O-based ligands, and these ligands can replace phosphine ligands in iron chemistry.

As the atmosphere on Earth changed following the Great Oxidation Event about 2.4 billion years ago, ferrous (+2) iron complexes became less stable and ferric (+3) complexes became predominant [2].

Iron in its ferric oxidation state typically forms complexes that are water-insoluble like hematite (Fe2O3) or magnetite (Fe3O4), especially under basic conditions when exposed to air. This propensity of iron complexes to precipitate can be a hindrance for catalysis, although, despite the fact that ferric complexes are mostly water-insoluble at biological pH, iron is still the most common transition metal in living organisms and is indispensable for the chemical processes of life—oxygen binding, electron transport, DNA synthesis, and cell proliferation, to name only a few.

Modern catalysis has been dominated by noble transition metals, such as palladium, platinum, ruthenium and iridium, and these metals have been used in a wide range of reactions. The main

Molecules 2020, 25, 1349; doi:10.3390/molecules25061349 www.mdpi.com/journal/molecules

(2)

Molecules 2020, 25, 1349 2 of 20

advantage that noble transition metals have over their base counterparts is their preference for undergoing two-electron processes. They do, however, have significant drawbacks: high cost, non-renewable supply and/or precarious toxicological and ecological properties. These factors may not pose too much of a problem for academic research, but have profound implications for industry and for the future of sustainable chemistry. It is for these reasons that attention has shifted towards base transition metals like iron, copper and nickel. Iron is particularly well suited as it is the second most abundant metal in the Earth’s crust after aluminum and is consequently attractive economically and ecologically. Despite these advantages of iron, organoiron catalysts tend to suffer from serious drawbacks such as difficult synthetic pathways, lack of robustness, poor atom economy and low activity or enantioselectivity. Although circumventing these limitations will be necessary for iron catalysis to reach its full potential, base metal catalysis will no doubt gain importance in the future and it is reasonable to think that base metals such as iron will eventually supplant the traditional dominance of noble transition metals as the field matures. In recent times, the area of iron catalysis has exploded and Beller in 2008 and Bolm in 2009 declared that the age of iron has begun [3,4]. An intriguing outlook on the future of homogeneous iron catalysis was published in 2016 by Fürstner [5].

This review focuses on the recent advancements in iron catalysis pertaining to organic chemistry from 2016 to February 2020. An excellent and comprehensive review from 2015 by Knölker on all the types of iron-catalyzed reactions discussed in this review in addition to others can be consulted for the interested reader [6]. Other more specialized reviews may be found in each respective subsection.

It should be mentioned that metathesis reactions are omitted from this review.

2. Iron in Organic Synthesis

Organoiron chemistry began in 1891 with the discovery of iron pentacarbonyl by Mond and Berthelot [7,8]. It was used sixty years later industrially in the Reppe process of hydroformylation of ethylene to form propionaldehyde and 1-propanol in basic solutions [9]. An important event was the discovery of ferrocene in 1951 [10], the structure of which was determined in 1952 [11,12] and led to the Nobel prize being awarded to Wilkinson and Fischer in 1973. The discovery of the Haber–Bosch process was an additional milestone in iron chemistry. The latter process uses an inorganic iron catalyst for the production of ammonia and sparked an agricultural revolution [13,14]. In modern organoiron chemistry, iron is used in a great number of diverse reactions, as will be apparent from this review, though perhaps, just as in the chemistry of life, its most ubiquitous role is in redox chemistry.

2.1. Addition Reactions

The first example of an iron-catalyzed racemization of alcohols was reported in 2016 with the use of an iron pincer catalyst [15]. Between 2016 and 2017, the groups of Bäckvall [16] and Rueping [17]

independently reported the dynamic kinetic resolution (DKR) of sec-alcohols using a combination of iron catalysis for racemization and a lipase for resolution, which demonstrates a useful combination of enzyme and transition metal catalysis (Scheme1). In one study, Knölkers complex (II) was used directly [17], and in the other study, a bench-stable precursor to Knölkers complex (II), iron tricarbonyl complex I, was used, which was activated through oxidative decarbonylation with TMANO to form coordinatively unsaturated iron complex I’ [16]. In the latter study, various benzylic and aliphatic esters could be produced in good to excellent yields with excellent ee. Two different enzymes, Candida antarctica lipase B (CalB/Novozyme 435) and Burkholderia cepasia (PS-C), could be used and the procedure could be reproduced on gram-scale. The group of Zhou also reported a related work using hexanoate as the acyl donor [18]. Rodriguez published a review on the synthesis, properties and reactivity of this interesting class of iron catalysts in 2015 [19].

(3)

Molecules 2020, 25, 1349 3 of 20

Molecules 2020, 25, x FOR PEER REVIEW 3 of 20

90

Scheme 1. Bäckvall’s and Rueping’s DKR of sec-alcohols.

91

The indole ring is a ubiquitous heterocyclic motif in natural products and methods for

92

constructing chiral polycyclic systems with indole skeletons has attracted considerable attention. In

93

2017, the group of Zhou reported the first intramolecular enantioselective cyclopropanation of

94

indoles that was catalyzed either by iron or copper in the presence of a chiral ligand (Scheme 2) [20].

95

Many functional groups were tolerated and various cyclopropanated indoles were prepared in high

96

to excellent yields and in almost all cases with excellent ee. The mechanism of the enhancement of

97

the enantioselectivity is currently unknown, although the R2 group was found to be important and

98

had to be different from hydrogen for the reaction to proceed.

99

100

Scheme 2. Zhou’s enantioselective cyclopropanation.

101

Hydroamination of alkenes is an atom-economic approach and the amines produced are some

102

of the most common functionalities found in fine chemicals and pharmaceuticals. Hydroamination

103

of terminal alkenes typically gives the Markovnikov product selectivity, but in 2019 the group of

104

Wang reported the first iron-catalyzed anti-Markovnikov addition of allylic alcohols [21]. For this

105

purpose, an iron-PNP pincer complex was used. The reaction proceeds through a

106

hydrogen-borrowing strategy where the iron complex temporarily activates the alcohol by

107

dehydrogenation to the α,β-unsaturated carbonyl compound. The latter compound reacts with an

108

amine to form an iminium ion, which undergoes conjugate additon at the β-position with another

109

amine followed by hydrolysis and reduction to give the product. Various amines were produced in

110

good yields with this method. Interestingly, hydroamidation could also be performed.

111

Scheme 1.Bäckvall’s and Rueping’s DKR of sec-alcohols.

The indole ring is a ubiquitous heterocyclic motif in natural products and methods for constructing chiral polycyclic systems with indole skeletons has attracted considerable attention. In 2017, the group of Zhou reported the first intramolecular enantioselective cyclopropanation of indoles that was catalyzed either by iron or copper in the presence of a chiral ligand (Scheme2) [20]. Many functional groups were tolerated and various cyclopropanated indoles were prepared in high to excellent yields and in almost all cases with excellent ee. The mechanism of the enhancement of the enantioselectivity is currently unknown, although the R2group was found to be important and had to be different from hydrogen for the reaction to proceed.

90

Scheme 1. Bäckvall’s and Rueping’s DKR of sec-alcohols.

91

The indole ring is a ubiquitous heterocyclic motif in natural products and methods for

92

constructing chiral polycyclic systems with indole skeletons has attracted considerable attention. In

93

2017, the group of Zhou reported the first intramolecular enantioselective cyclopropanation of

94

indoles that was catalyzed either by iron or copper in the presence of a chiral ligand (Scheme 2) [20].

95

Many functional groups were tolerated and various cyclopropanated indoles were prepared in high

96

to excellent yields and in almost all cases with excellent ee. The mechanism of the enhancement of

97

the enantioselectivity is currently unknown, although the R2 group was found to be important and

98

had to be different from hydrogen for the reaction to proceed.

99

100

Scheme 2. Zhou’s enantioselective cyclopropanation.

101

Hydroamination of alkenes is an atom-economic approach and the amines produced are some

102

of the most common functionalities found in fine chemicals and pharmaceuticals. Hydroamination

103

of terminal alkenes typically gives the Markovnikov product selectivity, but in 2019 the group of

104

Wang reported the first iron-catalyzed anti-Markovnikov addition of allylic alcohols [21]. For this

105

purpose, an iron-PNP pincer complex was used. The reaction proceeds through a

106

hydrogen-borrowing strategy where the iron complex temporarily activates the alcohol by

107

dehydrogenation to the α,β-unsaturated carbonyl compound. The latter compound reacts with an

108

amine to form an iminium ion, which undergoes conjugate additon at the β-position with another

109

amine followed by hydrolysis and reduction to give the product. Various amines were produced in

110

good yields with this method. Interestingly, hydroamidation could also be performed.

111

Scheme 2.Zhou’s enantioselective cyclopropanation.

Hydroamination of alkenes is an atom-economic approach and the amines produced are some of the most common functionalities found in fine chemicals and pharmaceuticals. Hydroamination of terminal alkenes typically gives the Markovnikov product selectivity, but in 2019 the group of Wang reported the first iron-catalyzed anti-Markovnikov addition of allylic alcohols [21]. For this purpose, an iron-PNP pincer complex was used. The reaction proceeds through a hydrogen-borrowing strategy where the iron complex temporarily activates the alcohol by dehydrogenation to the α,β-unsaturated carbonyl compound. The latter compound reacts with an amine to form an iminium ion, which undergoes conjugate additon at the β-position with another amine followed by hydrolysis and reduction to give the product. Various amines were produced in good yields with this method.

Interestingly, hydroamidation could also be performed (Scheme3).

(4)

Molecules 2020, 25, 1349 4 of 20

Molecules 2020, 25, x FOR PEER REVIEW 4 of 20

112

Scheme 3. Wang’s iron-catalyzed anti-Markovnikov hydroamination.

113

C-C and C-N bonds are important bonds in organic chemistry and one of the most effective

114

ways of creating these bonds simultaneously is through the carboamination of olefins. In 2017, the

115

group of Bao reported the diastereoselective construction of amines and disubstituted β-amino acids

116

through the carboamination of olefins (Scheme 4) [22]. Aliphatic acids were used as an alkyl source

117

and nitriles as a nitrogen source. The protocol could be performed on a gram-scale and various

118

carboamination products were obtained in good yields and excellent diastereoselectivity. The choice

119

of acid was found to have a strong effect on the diastereoselectivity. TsOH was used for the

120

carboamination of olefins, but for the carboamination of esters, binary and ternary acids had a more

121

positive effect over monoacids, with H2SO4 giving the best result. The addition of the nitrile group

122

was found, through Density Functional Theory (DFT) calculations, to be

123

diastereoselectivity-determining, and hyperconjugation was proposed to account for the

124

anti-selectivity.

125

126

Scheme 4. Bao’s carboamination.

127

Organosilicon compounds have significant chemical, physical and bioactive properties and an

128

example of these compounds is 1-amino-2-silylalkanes, which have, in recent times, emerged as

129

candidates for pharmaceutical development. Silicon-containing compounds are generally made

130

through hydrosylilation or dehydrogenative silylation, but in 2017 the group of Luo reported the

131

first iron-catalyzed synthesis of 1-amino-2-silylalkanes through the 1,2-difunctionalization of

132

styrenes and conjugated alkenes (Scheme 5) [23]. Di-tert-butyl-peroxide (DTBP) was used as an

133

oxidant in the reaction. Amines, amides and carbon nucleophiles could be employed and delivered

134

the corresponding products in mostly good yields. The reaction was proposed to proceed via a

135

silicon-centered radical from oxidative cleavage of the Si-H bond followed by addition across the

136

C=C bond and a N-H oxidative functionalization cascade.

137

138

Scheme 3.Wang’s iron-catalyzed anti-Markovnikov hydroamination.

C-C and C-N bonds are important bonds in organic chemistry and one of the most effective ways of creating these bonds simultaneously is through the carboamination of olefins. In 2017, the group of Bao reported the diastereoselective construction of amines and disubstituted β-amino acids through the carboamination of olefins (Scheme4) [22]. Aliphatic acids were used as an alkyl source and nitriles as a nitrogen source. The protocol could be performed on a gram-scale and various carboamination products were obtained in good yields and excellent diastereoselectivity. The choice of acid was found to have a strong effect on the diastereoselectivity. TsOH was used for the carboamination of olefins, but for the carboamination of esters, binary and ternary acids had a more positive effect over monoacids, with H2SO4giving the best result. The addition of the nitrile group was found, through Density Functional Theory (DFT) calculations, to be diastereoselectivity-determining, and hyperconjugation was proposed to account for the anti-selectivity.

Molecules 2020, 25, x FOR PEER REVIEW 4 of 20

112

Scheme 3. Wang’s iron-catalyzed anti-Markovnikov hydroamination.

113

C-C and C-N bonds are important bonds in organic chemistry and one of the most effective

114

ways of creating these bonds simultaneously is through the carboamination of olefins. In 2017, the

115

group of Bao reported the diastereoselective construction of amines and disubstituted β-amino acids

116

through the carboamination of olefins (Scheme 4) [22]. Aliphatic acids were used as an alkyl source

117

and nitriles as a nitrogen source. The protocol could be performed on a gram-scale and various

118

carboamination products were obtained in good yields and excellent diastereoselectivity. The choice

119

of acid was found to have a strong effect on the diastereoselectivity. TsOH was used for the

120

carboamination of olefins, but for the carboamination of esters, binary and ternary acids had a more

121

positive effect over monoacids, with H2SO4 giving the best result. The addition of the nitrile group

122

was found, through Density Functional Theory (DFT) calculations, to be

123

diastereoselectivity-determining, and hyperconjugation was proposed to account for the

124

anti-selectivity.

125

126

Scheme 4. Bao’s carboamination.

127

Organosilicon compounds have significant chemical, physical and bioactive properties and an

128

example of these compounds is 1-amino-2-silylalkanes, which have, in recent times, emerged as

129

candidates for pharmaceutical development. Silicon-containing compounds are generally made

130

through hydrosylilation or dehydrogenative silylation, but in 2017 the group of Luo reported the

131

first iron-catalyzed synthesis of 1-amino-2-silylalkanes through the 1,2-difunctionalization of

132

styrenes and conjugated alkenes (Scheme 5) [23]. Di-tert-butyl-peroxide (DTBP) was used as an

133

oxidant in the reaction. Amines, amides and carbon nucleophiles could be employed and delivered

134

the corresponding products in mostly good yields. The reaction was proposed to proceed via a

135

silicon-centered radical from oxidative cleavage of the Si-H bond followed by addition across the

136

C=C bond and a N-H oxidative functionalization cascade.

137

138

Scheme 4.Bao’s carboamination.

Organosilicon compounds have significant chemical, physical and bioactive properties and an example of these compounds is 1-amino-2-silylalkanes, which have, in recent times, emerged as candidates for pharmaceutical development. Silicon-containing compounds are generally made through hydrosylilation or dehydrogenative silylation, but in 2017 the group of Luo reported the first iron-catalyzed synthesis of 1-amino-2-silylalkanes through the 1,2-difunctionalization of styrenes and conjugated alkenes (Scheme5) [23]. Di-tert-butyl-peroxide (DTBP) was used as an oxidant in the reaction. Amines, amides and carbon nucleophiles could be employed and delivered the corresponding products in mostly good yields. The reaction was proposed to proceed via a silicon-centered radical from oxidative cleavage of the Si-H bond followed by addition across the C=C bond and a N-H oxidative functionalization cascade.

Molecules 2020, 25, x FOR PEER REVIEW 4 of 20

112

Scheme 3. Wang’s iron-catalyzed anti-Markovnikov hydroamination.

113

C-C and C-N bonds are important bonds in organic chemistry and one of the most effective

114

ways of creating these bonds simultaneously is through the carboamination of olefins. In 2017, the

115

group of Bao reported the diastereoselective construction of amines and disubstituted β-amino acids

116

through the carboamination of olefins (Scheme 4) [22]. Aliphatic acids were used as an alkyl source

117

and nitriles as a nitrogen source. The protocol could be performed on a gram-scale and various

118

carboamination products were obtained in good yields and excellent diastereoselectivity. The choice

119

of acid was found to have a strong effect on the diastereoselectivity. TsOH was used for the

120

carboamination of olefins, but for the carboamination of esters, binary and ternary acids had a more

121

positive effect over monoacids, with H2SO4 giving the best result. The addition of the nitrile group

122

was found, through Density Functional Theory (DFT) calculations, to be

123

diastereoselectivity-determining, and hyperconjugation was proposed to account for the

124

anti-selectivity.

125

126

Scheme 4. Bao’s carboamination.

127

Organosilicon compounds have significant chemical, physical and bioactive properties and an

128

example of these compounds is 1-amino-2-silylalkanes, which have, in recent times, emerged as

129

candidates for pharmaceutical development. Silicon-containing compounds are generally made

130

through hydrosylilation or dehydrogenative silylation, but in 2017 the group of Luo reported the

131

first iron-catalyzed synthesis of 1-amino-2-silylalkanes through the 1,2-difunctionalization of

132

styrenes and conjugated alkenes (Scheme 5) [23]. Di-tert-butyl-peroxide (DTBP) was used as an

133

oxidant in the reaction. Amines, amides and carbon nucleophiles could be employed and delivered

134

the corresponding products in mostly good yields. The reaction was proposed to proceed via a

135

silicon-centered radical from oxidative cleavage of the Si-H bond followed by addition across the

136

C=C bond and a N-H oxidative functionalization cascade.

137

138

Scheme 5.Luo’s silylation.

(5)

Molecules 2020, 25, 1349 5 of 20

2.2. C-H Bond Activation

The first example of a C-H activation was a Friedel–Crafts reaction reported by Dimroth in 1902, where benzene reacted with mercury (II) acetate to give phenylmercury (II) acetate [24].

Later, in 1955, Murahashi reported the cobalt-catalyzed chelation-assisted C-H functionalization of (E)-N-1-diphenylmethaneimine to 2-phenylisoindolin-1-one [25]. A great advance in the field occurred in 1966 when Shilov reported that K2PtCl4could induce isotope scrambling between methane and heavy water [26,27]. Shilov’s discovery led to the so called “Shilov system”, which remains to this day as one of the few catalytic systems that can accomplish selective alkene functionalizations under mild conditions. In 2008, the synthetic power of C-H activation was expanded to include organoiron catalysis by Nakamura in his arylation of benzoquinolines (Scheme6) [28]. An excellent review on the subject of iron in C-H activation reactions by Nakamura was published in 2017 [29] and a review on oxidative C-H activation was published by Li in 2014 [30].

Molecules 2020, 25, x FOR PEER REVIEW 5 of 20

Scheme 5. Luo’s silylation.

139

2.2. C-H Bond Activation

140

The first example of a C-H activation was a Friedel–Crafts reaction reported by Dimroth in

141

1902, where benzene reacted with mercury (II) acetate to give phenylmercury (II) acetate [24]. Later,

142

in 1955, Murahashi reported the cobalt-catalyzed chelation-assisted C-H functionalization of

143

(E)-N-1-diphenylmethaneimine to 2-phenylisoindolin-1-one [25]. A great advance in the field

144

occurred in 1966 when Shilov reported that K2PtCl4 could induce isotope scrambling between

145

methane and heavy water [26,27]. Shilov’s discovery led to the so called “Shilov system”, which

146

remains to this day as one of the few catalytic systems that can accomplish selective alkene

147

functionalizations under mild conditions. In 2008, the synthetic power of C-H activation was

148

expanded to include organoiron catalysis by Nakamura in his arylation of benzoquinolines (Scheme

149

6) [28]. An excellent review on the subject of iron in C-H activation reactions by Nakamura was

150

published in 2017 [29] and a review on oxidative C-H activation was published by Li in 2014 [30].

151

152

Scheme 6. Nakamura’s C-H activation.

153

The group of Arnold has in the past used engineered cytochrome P450, which is a type of enzyme

154

that uses a heme cofactor, to enantioselectively α-hydroxylate arylacetic acid derivatives via C-H

155

activation [31]. In 2017, they reported the directed evolution of cytochrome P450 monooxygenase, for

156

enantioselective C-H activation to give C-N bonds (Scheme 7) [32]. It uses a variant of P411 based on

157

the P450 monooxygenase which has an axial serine ligand on the haem iron instead of the natural

158

cysteine. The method utilizes a tosyl azide as a nitrene source which generates an iron nitrenoid that

159

subsequently reacts with an alkane to deliver the C-H amination product. The P411 variant has a

160

turnover number (TON) of 1300, which is considerably higher than the best reported, to our

161

knowledge, for traditional chiral transition metal complexes, which is a chiral manganese porphyrin

162

with a turnover number of 85 [33]. A variety of benzylic tosylamines could be produced with excellent

163

ees.

164

165

Scheme 7. Arnold’s C-H amination.

166

In 2019, Arnold and coworkers extended their methodology using another variant of P411 in

167

C-H alkylation using diazoesters (Scheme 8) [34]. The diazo substrate scope could be extended

168

beyond ester-based reagents to Weinreb amides and diazoketones and gave the corresponding

169

products with excellent ees and with total turnover numbers (TTN) of up to 2330. These studies

170

together show the potential for generating C-H alkylation enzymes that can emulate the scope and

171

selectivity of Natures C-H oxygenation catalysts.

172

Scheme 6.Nakamura’s C-H activation.

The group of Arnold has in the past used engineered cytochrome P450, which is a type of enzyme that uses a heme cofactor, to enantioselectively α-hydroxylate arylacetic acid derivatives via C-H activation [31]. In 2017, they reported the directed evolution of cytochrome P450 monooxygenase, for enantioselective C-H activation to give C-N bonds (Scheme7) [32]. It uses a variant of P411 based on the P450 monooxygenase which has an axial serine ligand on the haem iron instead of the natural cysteine. The method utilizes a tosyl azide as a nitrene source which generates an iron nitrenoid that subsequently reacts with an alkane to deliver the C-H amination product. The P411 variant has a turnover number (TON) of 1300, which is considerably higher than the best reported, to our knowledge, for traditional chiral transition metal complexes, which is a chiral manganese porphyrin with a turnover number of 85 [33]. A variety of benzylic tosylamines could be produced with excellent ees.

Scheme 5. Luo’s silylation.

139

2.2. C-H Bond Activation

140

The first example of a C-H activation was a Friedel–Crafts reaction reported by Dimroth in

141

1902, where benzene reacted with mercury (II) acetate to give phenylmercury (II) acetate [24]. Later,

142

in 1955, Murahashi reported the cobalt-catalyzed chelation-assisted C-H functionalization of

143

(E)-N-1-diphenylmethaneimine to 2-phenylisoindolin-1-one [25]. A great advance in the field

144

occurred in 1966 when Shilov reported that K2PtCl4 could induce isotope scrambling between

145

methane and heavy water [26,27]. Shilov’s discovery led to the so called “Shilov system”, which

146

remains to this day as one of the few catalytic systems that can accomplish selective alkene

147

functionalizations under mild conditions. In 2008, the synthetic power of C-H activation was

148

expanded to include organoiron catalysis by Nakamura in his arylation of benzoquinolines (Scheme

149

6) [28]. An excellent review on the subject of iron in C-H activation reactions by Nakamura was

150

published in 2017 [29] and a review on oxidative C-H activation was published by Li in 2014 [30].

151

152

Scheme 6. Nakamura’s C-H activation.

153

The group of Arnold has in the past used engineered cytochrome P450, which is a type of enzyme

154

that uses a heme cofactor, to enantioselectively α-hydroxylate arylacetic acid derivatives via C-H

155

activation [31]. In 2017, they reported the directed evolution of cytochrome P450 monooxygenase, for

156

enantioselective C-H activation to give C-N bonds (Scheme 7) [32]. It uses a variant of P411 based on

157

the P450 monooxygenase which has an axial serine ligand on the haem iron instead of the natural

158

cysteine. The method utilizes a tosyl azide as a nitrene source which generates an iron nitrenoid that

159

subsequently reacts with an alkane to deliver the C-H amination product. The P411 variant has a

160

turnover number (TON) of 1300, which is considerably higher than the best reported, to our

161

knowledge, for traditional chiral transition metal complexes, which is a chiral manganese porphyrin

162

with a turnover number of 85 [33]. A variety of benzylic tosylamines could be produced with excellent

163

ees.

164

165

Scheme 7. Arnold’s C-H amination.

166

In 2019, Arnold and coworkers extended their methodology using another variant of P411 in

167

C-H alkylation using diazoesters (Scheme 8) [34]. The diazo substrate scope could be extended

168

beyond ester-based reagents to Weinreb amides and diazoketones and gave the corresponding

169

products with excellent ees and with total turnover numbers (TTN) of up to 2330. These studies

170

together show the potential for generating C-H alkylation enzymes that can emulate the scope and

171

selectivity of Natures C-H oxygenation catalysts.

172

Scheme 7.Arnold’s C-H amination.

In 2019, Arnold and coworkers extended their methodology using another variant of P411 in C-H alkylation using diazoesters (Scheme8) [34]. The diazo substrate scope could be extended beyond ester-based reagents to Weinreb amides and diazoketones and gave the corresponding products with excellent ees and with total turnover numbers (TTN) of up to 2330. These studies together show the potential for generating C-H alkylation enzymes that can emulate the scope and selectivity of Natures C-H oxygenation catalysts.

(6)

Molecules 2020, 25, 1349 6 of 20

Molecules 2020, 25, x FOR PEER REVIEW 6 of 20

173

Scheme 8. Arnold’s sp3 C-H activation.

174

C(sp3)-H alkylation via an isoelectronic iron carbene intermediate was first reported in 2017 by

175

the group of White using an iron phthalocyanine (Scheme 9) [35]. Iron carbenes generally prefer

176

cyclopropanation over C-H oxidation, but, in this case, allylic and benzylic C(sp3)-H bonds could be

177

alkylated with a broad scope. Mechanistic studies indicated that an electrophilic iron carbene was

178

mediating homolytic C-H cleavage followed by recombination with the resulting alkyl radical to

179

form the new C-C bond. The C-H cleavage was found to be partially rate determining.

180

181

Scheme 9. White’s isoelectronic carbene C(sp3)-H oxidation.

182

In 2018, the group of Ackermann reported on an allene annulation through an iron-catalyzed

183

C-H/N-H/C-O/C-H functionalization sequence (Scheme 10) [36]. The mechanism was shown to

184

involve an unprecedented 1,4-iron migration C-H activation manifold. Alkyl chlorides were

185

tolerated under these reaction conditions, with no cross-coupling being observed. Various

186

dihydroisoquinolones could be produced through the use of this method in excellent yields and the

187

modular nature of the triazole group allowed for the synthesis of exo-methylene isoquinolones as

188

well.

189

190

Scheme 10. Ackermann’s allene annulation.

191

In late 2019, the group of Wang demonstrated an iron-catalyzed α-C-H functionalization of

192

π-bonds in the hydroxyalkylation of alkynes and olefins (Scheme 11) [37]. Propargylic and allylic

193

C-H bonds were functionalized with this method and a wide variety of homopropargylic and

194

homoallylic alcohols could be produced in excellent yields, although with modest stereoselectivity.

195

The key to the success of this approach is the fact that coordination of the iron catalyst to the

196

unsaturated bond is known to lower the pKa of a propargylic or allylic proton from ≈38 and ≈43,

197

respectively, to <10 [38]. An (α-allenyl)iron or (π-allyl)iron complex for propargylic or allylic

198

complexes, respectively, is then formed in the presence of a base, which is utilized as the coupling

199

partner.

200

Scheme 8.Arnold’s sp3C-H activation.

C(sp3)-H alkylation via an isoelectronic iron carbene intermediate was first reported in 2017 by the group of White using an iron phthalocyanine (Scheme9) [35]. Iron carbenes generally prefer cyclopropanation over C-H oxidation, but, in this case, allylic and benzylic C(sp3)-H bonds could be alkylated with a broad scope. Mechanistic studies indicated that an electrophilic iron carbene was mediating homolytic C-H cleavage followed by recombination with the resulting alkyl radical to form the new C-C bond. The C-H cleavage was found to be partially rate determining.

Molecules 2020, 25, x FOR PEER REVIEW 6 of 20

173

Scheme 8. Arnold’s sp3 C-H activation.

174

C(sp3)-H alkylation via an isoelectronic iron carbene intermediate was first reported in 2017 by

175

the group of White using an iron phthalocyanine (Scheme 9) [35]. Iron carbenes generally prefer

176

cyclopropanation over C-H oxidation, but, in this case, allylic and benzylic C(sp3)-H bonds could be

177

alkylated with a broad scope. Mechanistic studies indicated that an electrophilic iron carbene was

178

mediating homolytic C-H cleavage followed by recombination with the resulting alkyl radical to

179

form the new C-C bond. The C-H cleavage was found to be partially rate determining.

180

181

Scheme 9. White’s isoelectronic carbene C(sp3)-H oxidation.

182

In 2018, the group of Ackermann reported on an allene annulation through an iron-catalyzed

183

C-H/N-H/C-O/C-H functionalization sequence (Scheme 10) [36]. The mechanism was shown to

184

involve an unprecedented 1,4-iron migration C-H activation manifold. Alkyl chlorides were

185

tolerated under these reaction conditions, with no cross-coupling being observed. Various

186

dihydroisoquinolones could be produced through the use of this method in excellent yields and the

187

modular nature of the triazole group allowed for the synthesis of exo-methylene isoquinolones as

188

well.

189

190

Scheme 10. Ackermann’s allene annulation.

191

In late 2019, the group of Wang demonstrated an iron-catalyzed α-C-H functionalization of

192

π-bonds in the hydroxyalkylation of alkynes and olefins (Scheme 11) [37]. Propargylic and allylic

193

C-H bonds were functionalized with this method and a wide variety of homopropargylic and

194

homoallylic alcohols could be produced in excellent yields, although with modest stereoselectivity.

195

The key to the success of this approach is the fact that coordination of the iron catalyst to the

196

unsaturated bond is known to lower the pKa of a propargylic or allylic proton from ≈38 and ≈43,

197

respectively, to <10 [38]. An (α-allenyl)iron or (π-allyl)iron complex for propargylic or allylic

198

complexes, respectively, is then formed in the presence of a base, which is utilized as the coupling

199

partner.

200

Scheme 9.White’s isoelectronic carbene C(sp3)-H oxidation.

In 2018, the group of Ackermann reported on an allene annulation through an iron-catalyzed C-H/N-H/C-O/C-H functionalization sequence (Scheme10) [36]. The mechanism was shown to involve an unprecedented 1,4-iron migration C-H activation manifold. Alkyl chlorides were tolerated under these reaction conditions, with no cross-coupling being observed. Various dihydroisoquinolones could be produced through the use of this method in excellent yields and the modular nature of the triazole group allowed for the synthesis of exo-methylene isoquinolones as well.

Molecules 2020, 25, x FOR PEER REVIEW 6 of 20

173

Scheme 8. Arnold’s sp3 C-H activation.

174

C(sp3)-H alkylation via an isoelectronic iron carbene intermediate was first reported in 2017 by

175

the group of White using an iron phthalocyanine (Scheme 9) [35]. Iron carbenes generally prefer

176

cyclopropanation over C-H oxidation, but, in this case, allylic and benzylic C(sp3)-H bonds could be

177

alkylated with a broad scope. Mechanistic studies indicated that an electrophilic iron carbene was

178

mediating homolytic C-H cleavage followed by recombination with the resulting alkyl radical to

179

form the new C-C bond. The C-H cleavage was found to be partially rate determining.

180

181

Scheme 9. White’s isoelectronic carbene C(sp3)-H oxidation.

182

In 2018, the group of Ackermann reported on an allene annulation through an iron-catalyzed

183

C-H/N-H/C-O/C-H functionalization sequence (Scheme 10) [36]. The mechanism was shown to

184

involve an unprecedented 1,4-iron migration C-H activation manifold. Alkyl chlorides were

185

tolerated under these reaction conditions, with no cross-coupling being observed. Various

186

dihydroisoquinolones could be produced through the use of this method in excellent yields and the

187

modular nature of the triazole group allowed for the synthesis of exo-methylene isoquinolones as

188

well.

189

190

Scheme 10. Ackermann’s allene annulation.

191

In late 2019, the group of Wang demonstrated an iron-catalyzed α-C-H functionalization of

192

π-bonds in the hydroxyalkylation of alkynes and olefins (Scheme 11) [37]. Propargylic and allylic

193

C-H bonds were functionalized with this method and a wide variety of homopropargylic and

194

homoallylic alcohols could be produced in excellent yields, although with modest stereoselectivity.

195

The key to the success of this approach is the fact that coordination of the iron catalyst to the

196

unsaturated bond is known to lower the pKa of a propargylic or allylic proton from ≈38 and ≈43,

197

respectively, to <10 [38]. An (α-allenyl)iron or (π-allyl)iron complex for propargylic or allylic

198

complexes, respectively, is then formed in the presence of a base, which is utilized as the coupling

199

partner.

200

Scheme 10.Ackermann’s allene annulation.

In late 2019, the group of Wang demonstrated an iron-catalyzed α-C-H functionalization of π-bonds in the hydroxyalkylation of alkynes and olefins (Scheme11) [37]. Propargylic and allylic C-H bonds were functionalized with this method and a wide variety of homopropargylic and homoallylic alcohols could be produced in excellent yields, although with modest stereoselectivity. The key to the success of this approach is the fact that coordination of the iron catalyst to the unsaturated bond is known to lower the pKa of a propargylic or allylic proton from ≈38 and ≈43, respectively, to<10 [38].

An (α-allenyl)iron or (π-allyl)iron complex for propargylic or allylic complexes, respectively, is then formed in the presence of a base, which is utilized as the coupling partner.

(7)

Molecules 2020, 25, x FOR PEER REVIEW 7 of 20

201

Scheme 11. Wang’s α-C-H functionalization.

202

In 2018, the group of Liu reported the unprecedented iron(II)-catalyzed fluorination of C(sp3)-H

203

bonds using alkoxyl radicals (Scheme 12) [39]. The procedure was applied to a wide range of

204

substrates and it was found that a range of functional groups were tolerated, including halide and

205

hydroxyl groups. N-fluorobenzenesulfonamide (NFSI) was used as the fluoride source and the

206

substrate scope could be extended from fluorination to chlorination, amination and alkylation. The

207

authors also demonstrated a one-pot application of their protocol starting from a simple alkane.

208

209

Scheme 12. Liu’s C-H fluorination.

210

2.3. Cross-Coupling Reactions

211

Transition-metal-catalyzed cross coupling protocols have become an important tool in the

212

organic chemist‘s arsenal. This area has been important in chemistry for about five decades, and in

213

2010 it received formal recognition when Richard Heck, Akira Suzuki and Ei-ichi Negishi received

214

the Nobel prize for palladium-catalyzed cross-couplings in organic synthesis. Although a powerful

215

technique, its applications have been dominated by the use of expensive palladium- and

216

nickel-based catalysts, which are often toxic. The most common types of cross coupling reactions

217

using iron are those involving Grignard reagents as the transmetalating nucleophile. The first

218

example of an alkenylation of alkyl Grignard reagents with organic halides using iron(III) chloride

219

was reported in 1971 by Kochi and Tamura (Scheme 13) [40]. A review on the subject of

220

iron-catalyzed cross-coupling reactions with a focus on mechanistic studies was published in 2016

221

by Byers [41]. A more focused review on the use of iron-catalyzed cross-coupling for the synthesis of

222

pharmaceuticals was released in 2018 by Szostak [42].

223

224

Scheme 13. Kochi’s and Tamura’s original alkenylation.

225

In 2016, Bäckvall and coworkers reported the coupling of propargyl carboxylates and Grignard

226

reagents using the environmentally benign Fe(acac)3 to synthesize substituted allenes and protected

227

α-allenols (Scheme 14) [43,44]. The mild reaction conditions tolerate a broad range of functional

228

groups (silyl ethers, carbamates and acetals) and could be applied to more complex molecules such

229

as steroids. Tri and tetra substituted allenes were obtained in excellent yields, whereas the yield was

230

found to drop for less substituted allenes. A variety of alkyl and aryl Grignard reagents could be

231

applied and it was demonstrated that the protocol can be readily performed on a gram-scale.

232

Scheme 11.Wang’s α-C-H functionalization.

In 2018, the group of Liu reported the unprecedented iron(II)-catalyzed fluorination of C(sp3)-H bonds using alkoxyl radicals (Scheme12) [39]. The procedure was applied to a wide range of substrates and it was found that a range of functional groups were tolerated, including halide and hydroxyl groups. N-fluorobenzenesulfonamide (NFSI) was used as the fluoride source and the substrate scope could be extended from fluorination to chlorination, amination and alkylation. The authors also demonstrated a one-pot application of their protocol starting from a simple alkane.

Molecules 2020, 25, x FOR PEER REVIEW 7 of 20

201

Scheme 11. Wang’s α-C-H functionalization.

202

In 2018, the group of Liu reported the unprecedented iron(II)-catalyzed fluorination of C(sp3)-H

203

bonds using alkoxyl radicals (Scheme 12) [39]. The procedure was applied to a wide range of

204

substrates and it was found that a range of functional groups were tolerated, including halide and

205

hydroxyl groups. N-fluorobenzenesulfonamide (NFSI) was used as the fluoride source and the

206

substrate scope could be extended from fluorination to chlorination, amination and alkylation. The

207

authors also demonstrated a one-pot application of their protocol starting from a simple alkane.

208

209

Scheme 12. Liu’s C-H fluorination.

210

2.3. Cross-Coupling Reactions

211

Transition-metal-catalyzed cross coupling protocols have become an important tool in the

212

organic chemist‘s arsenal. This area has been important in chemistry for about five decades, and in

213

2010 it received formal recognition when Richard Heck, Akira Suzuki and Ei-ichi Negishi received

214

the Nobel prize for palladium-catalyzed cross-couplings in organic synthesis. Although a powerful

215

technique, its applications have been dominated by the use of expensive palladium- and

216

nickel-based catalysts, which are often toxic. The most common types of cross coupling reactions

217

using iron are those involving Grignard reagents as the transmetalating nucleophile. The first

218

example of an alkenylation of alkyl Grignard reagents with organic halides using iron(III) chloride

219

was reported in 1971 by Kochi and Tamura (Scheme 13) [40]. A review on the subject of

220

iron-catalyzed cross-coupling reactions with a focus on mechanistic studies was published in 2016

221

by Byers [41]. A more focused review on the use of iron-catalyzed cross-coupling for the synthesis of

222

pharmaceuticals was released in 2018 by Szostak [42].

223

224

Scheme 13. Kochi’s and Tamura’s original alkenylation.

225

In 2016, Bäckvall and coworkers reported the coupling of propargyl carboxylates and Grignard

226

reagents using the environmentally benign Fe(acac)3 to synthesize substituted allenes and protected

227

α-allenols (Scheme 14) [43,44]. The mild reaction conditions tolerate a broad range of functional

228

groups (silyl ethers, carbamates and acetals) and could be applied to more complex molecules such

229

as steroids. Tri and tetra substituted allenes were obtained in excellent yields, whereas the yield was

230

found to drop for less substituted allenes. A variety of alkyl and aryl Grignard reagents could be

231

applied and it was demonstrated that the protocol can be readily performed on a gram-scale.

232

Scheme 12.Liu’s C-H fluorination.

2.3. Cross-Coupling Reactions

Transition-metal-catalyzed cross coupling protocols have become an important tool in the organic chemist‘s arsenal. This area has been important in chemistry for about five decades, and in 2010 it received formal recognition when Richard Heck, Akira Suzuki and Ei-ichi Negishi received the Nobel prize for palladium-catalyzed cross-couplings in organic synthesis. Although a powerful technique, its applications have been dominated by the use of expensive palladium- and nickel-based catalysts, which are often toxic. The most common types of cross coupling reactions using iron are those involving Grignard reagents as the transmetalating nucleophile. The first example of an alkenylation of alkyl Grignard reagents with organic halides using iron(III) chloride was reported in 1971 by Kochi and Tamura (Scheme13) [40]. A review on the subject of iron-catalyzed cross-coupling reactions with a focus on mechanistic studies was published in 2016 by Byers [41]. A more focused review on the use of iron-catalyzed cross-coupling for the synthesis of pharmaceuticals was released in 2018 by Szostak [42].

Molecules 2020, 25, x FOR PEER REVIEW 7 of 20

201

Scheme 11. Wang’s α-C-H functionalization.

202

In 2018, the group of Liu reported the unprecedented iron(II)-catalyzed fluorination of C(sp3)-H

203

bonds using alkoxyl radicals (Scheme 12) [39]. The procedure was applied to a wide range of

204

substrates and it was found that a range of functional groups were tolerated, including halide and

205

hydroxyl groups. N-fluorobenzenesulfonamide (NFSI) was used as the fluoride source and the

206

substrate scope could be extended from fluorination to chlorination, amination and alkylation. The

207

authors also demonstrated a one-pot application of their protocol starting from a simple alkane.

208

209

Scheme 12. Liu’s C-H fluorination.

210

2.3. Cross-Coupling Reactions

211

Transition-metal-catalyzed cross coupling protocols have become an important tool in the

212

organic chemist‘s arsenal. This area has been important in chemistry for about five decades, and in

213

2010 it received formal recognition when Richard Heck, Akira Suzuki and Ei-ichi Negishi received

214

the Nobel prize for palladium-catalyzed cross-couplings in organic synthesis. Although a powerful

215

technique, its applications have been dominated by the use of expensive palladium- and

216

nickel-based catalysts, which are often toxic. The most common types of cross coupling reactions

217

using iron are those involving Grignard reagents as the transmetalating nucleophile. The first

218

example of an alkenylation of alkyl Grignard reagents with organic halides using iron(III) chloride

219

was reported in 1971 by Kochi and Tamura (Scheme 13) [40]. A review on the subject of

220

iron-catalyzed cross-coupling reactions with a focus on mechanistic studies was published in 2016

221

by Byers [41]. A more focused review on the use of iron-catalyzed cross-coupling for the synthesis of

222

pharmaceuticals was released in 2018 by Szostak [42].

223

224

Scheme 13. Kochi’s and Tamura’s original alkenylation.

225

In 2016, Bäckvall and coworkers reported the coupling of propargyl carboxylates and Grignard

226

reagents using the environmentally benign Fe(acac)3 to synthesize substituted allenes and protected

227

α-allenols (Scheme 14) [43,44]. The mild reaction conditions tolerate a broad range of functional

228

groups (silyl ethers, carbamates and acetals) and could be applied to more complex molecules such

229

as steroids. Tri and tetra substituted allenes were obtained in excellent yields, whereas the yield was

230

found to drop for less substituted allenes. A variety of alkyl and aryl Grignard reagents could be

231

applied and it was demonstrated that the protocol can be readily performed on a gram-scale.

232

Scheme 13.Kochi’s and Tamura’s original alkenylation.

In 2016, Bäckvall and coworkers reported the coupling of propargyl carboxylates and Grignard reagents using the environmentally benign Fe(acac)3to synthesize substituted allenes and protected α-allenols (Scheme14) [43,44]. The mild reaction conditions tolerate a broad range of functional groups (silyl ethers, carbamates and acetals) and could be applied to more complex molecules such as steroids.

Tri and tetra substituted allenes were obtained in excellent yields, whereas the yield was found to drop for less substituted allenes. A variety of alkyl and aryl Grignard reagents could be applied and it was demonstrated that the protocol can be readily performed on a gram-scale.

(8)

Molecules 2020, 25, 1349 8 of 20

Molecules 2020, 25, x FOR PEER REVIEW 8 of 20

233

Scheme 14. Bäckvall’s synthesis of substituted allenes and protected α-allenols from carboxylates.

234

In 2016, the group of Frantz reported on a highly stereoselective iron-catalyzed cross coupling

235

using FeCl3 to couple Grignard reagents and enol carbamates (Scheme 15) [45]. Many functional

236

groups, such as ethers, silanes, primary bromides, alkynes and alkenes, were tolerated. In almost all

237

cases, the yield and E/Z selectivity was excellent, with (E)-carbamates leading to (E)-acrylates and

238

(Z)-carbamates leading to (Z)-acrylates. This study constitutes the only example so far of an

239

iron-catalyzed cross-coupling, where an oxygen-based electrophile is favored over a vinylic halide (a

240

Cl group at R2 in Scheme 15).

241

242

Scheme 15. Frantz’ stereoselective synthesis of acrylates.

243

Aryl C-glycosides are interesting pharmaceutical candidates because of their biological

244

activities and resistance to metabolic degradation. In 2017, the group of Nakamura developed a

245

highly diastereoselective iron-catalyzed cross-coupling of glycosyl halides and aryl metal reagents to

246

form these compounds using FeCl2 in conjunction with a SciOPP ligand (Scheme 16) [46]. A variety

247

of aryl, heteroaryl and vinyl metal reagents based on magnesium, zinc, boron and aluminium could

248

be applied. The reaction was found to proceed through the generation and stereoselective trapping

249

of glycosyl radical intermediates and represents a rare example of a highly stereoselective

250

carbon-carbon bond formation based on iron catalysis.

251

252

Scheme 16. Nakamura’s diastereoselective synthesis of aryl C-glycosides using (Sciopp) FeCl2.

253

Scheme 14.Bäckvall’s synthesis of substituted allenes and protected α-allenols from carboxylates.

In 2016, the group of Frantz reported on a highly stereoselective iron-catalyzed cross coupling using FeCl3to couple Grignard reagents and enol carbamates (Scheme15) [45]. Many functional groups, such as ethers, silanes, primary bromides, alkynes and alkenes, were tolerated. In almost all cases, the yield and E/Z selectivity was excellent, with (E)-carbamates leading to (E)-acrylates and (Z)-carbamates leading to (Z)-acrylates. This study constitutes the only example so far of an iron-catalyzed cross-coupling, where an oxygen-based electrophile is favored over a vinylic halide (a Cl group at R2in Scheme15).

Molecules 2020, 25, x FOR PEER REVIEW 8 of 20

233

Scheme 14. Bäckvall’s synthesis of substituted allenes and protected α-allenols from carboxylates.

234

In 2016, the group of Frantz reported on a highly stereoselective iron-catalyzed cross coupling

235

using FeCl3 to couple Grignard reagents and enol carbamates (Scheme 15) [45]. Many functional

236

groups, such as ethers, silanes, primary bromides, alkynes and alkenes, were tolerated. In almost all

237

cases, the yield and E/Z selectivity was excellent, with (E)-carbamates leading to (E)-acrylates and

238

(Z)-carbamates leading to (Z)-acrylates. This study constitutes the only example so far of an

239

iron-catalyzed cross-coupling, where an oxygen-based electrophile is favored over a vinylic halide (a

240

Cl group at R2 in Scheme 15).

241

242

Scheme 15. Frantz’ stereoselective synthesis of acrylates.

243

Aryl C-glycosides are interesting pharmaceutical candidates because of their biological

244

activities and resistance to metabolic degradation. In 2017, the group of Nakamura developed a

245

highly diastereoselective iron-catalyzed cross-coupling of glycosyl halides and aryl metal reagents to

246

form these compounds using FeCl2 in conjunction with a SciOPP ligand (Scheme 16) [46]. A variety

247

of aryl, heteroaryl and vinyl metal reagents based on magnesium, zinc, boron and aluminium could

248

be applied. The reaction was found to proceed through the generation and stereoselective trapping

249

of glycosyl radical intermediates and represents a rare example of a highly stereoselective

250

carbon-carbon bond formation based on iron catalysis.

251

252

Scheme 16. Nakamura’s diastereoselective synthesis of aryl C-glycosides using (Sciopp) FeCl2.

253

Scheme 15.Frantz’ stereoselective synthesis of acrylates.

Aryl C-glycosides are interesting pharmaceutical candidates because of their biological activities and resistance to metabolic degradation. In 2017, the group of Nakamura developed a highly diastereoselective iron-catalyzed cross-coupling of glycosyl halides and aryl metal reagents to form these compounds using FeCl2in conjunction with a SciOPP ligand (Scheme16) [46]. A variety of aryl, heteroaryl and vinyl metal reagents based on magnesium, zinc, boron and aluminium could be applied.

The reaction was found to proceed through the generation and stereoselective trapping of glycosyl radical intermediates and represents a rare example of a highly stereoselective carbon-carbon bond formation based on iron catalysis.

Molecules 2020, 25, x FOR PEER REVIEW 8 of 20

233

Scheme 14. Bäckvall’s synthesis of substituted allenes and protected α-allenols from carboxylates.

234

In 2016, the group of Frantz reported on a highly stereoselective iron-catalyzed cross coupling

235

using FeCl3 to couple Grignard reagents and enol carbamates (Scheme 15) [45]. Many functional

236

groups, such as ethers, silanes, primary bromides, alkynes and alkenes, were tolerated. In almost all

237

cases, the yield and E/Z selectivity was excellent, with (E)-carbamates leading to (E)-acrylates and

238

(Z)-carbamates leading to (Z)-acrylates. This study constitutes the only example so far of an

239

iron-catalyzed cross-coupling, where an oxygen-based electrophile is favored over a vinylic halide (a

240

Cl group at R2 in Scheme 15).

241

242

Scheme 15. Frantz’ stereoselective synthesis of acrylates.

243

Aryl C-glycosides are interesting pharmaceutical candidates because of their biological

244

activities and resistance to metabolic degradation. In 2017, the group of Nakamura developed a

245

highly diastereoselective iron-catalyzed cross-coupling of glycosyl halides and aryl metal reagents to

246

form these compounds using FeCl2 in conjunction with a SciOPP ligand (Scheme 16) [46]. A variety

247

of aryl, heteroaryl and vinyl metal reagents based on magnesium, zinc, boron and aluminium could

248

be applied. The reaction was found to proceed through the generation and stereoselective trapping

249

of glycosyl radical intermediates and represents a rare example of a highly stereoselective

250

carbon-carbon bond formation based on iron catalysis.

251

252

Scheme 16. Nakamura’s diastereoselective synthesis of aryl C-glycosides using (Sciopp) FeCl2.

253

Scheme 16.Nakamura’s diastereoselective synthesis of aryl C-glycosides using (Sciopp) FeCl2.

References

Related documents

The increasing availability of data and attention to services has increased the understanding of the contribution of services to innovation and productivity in

Syftet eller förväntan med denna rapport är inte heller att kunna ”mäta” effekter kvantita- tivt, utan att med huvudsakligt fokus på output och resultat i eller från

Generella styrmedel kan ha varit mindre verksamma än man har trott De generella styrmedlen, till skillnad från de specifika styrmedlen, har kommit att användas i större

I regleringsbrevet för 2014 uppdrog Regeringen åt Tillväxtanalys att ”föreslå mätmetoder och indikatorer som kan användas vid utvärdering av de samhällsekonomiska effekterna av

a) Inom den regionala utvecklingen betonas allt oftare betydelsen av de kvalitativa faktorerna och kunnandet. En kvalitativ faktor är samarbetet mellan de olika

Parallellmarknader innebär dock inte en drivkraft för en grön omställning Ökad andel direktförsäljning räddar många lokala producenter och kan tyckas utgöra en drivkraft

Närmare 90 procent av de statliga medlen (intäkter och utgifter) för näringslivets klimatomställning går till generella styrmedel, det vill säga styrmedel som påverkar

• Utbildningsnivåerna i Sveriges FA-regioner varierar kraftigt. I Stockholm har 46 procent av de sysselsatta eftergymnasial utbildning, medan samma andel i Dorotea endast